35
Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov * Amnon Ta-Shma Abstract The design of new quantum algorithms has proven to be an extremely difficult task. This paper considers a different approach to the problem. We systematically study ’quantum state generation’, namely, which super- positions can be efficiently generated. We first show that all problems in Statistical Zero Knowledge (SZK), a class which contains many languages that are natural candidates for BQP, can be reduced to an instance of quan- tum state generation. This was known before for graph isomorphism, but we give a general recipe for all problems in SZK. We demonstrate the re- duction from the problem to its quantum state generation version for three examples: Discrete log, quadratic residuosity and a gap version of closest vector in a lattice. We then develop tools for quantum state generation. For this task, we define the framework of ’adiabatic quantum state generation’ which uses the language of ground states, spectral gaps and Hamiltonians instead of the standard unitary gate language. This language stems from the recently suggested adiabatic computation model [20] and seems to be especially tai- lored for the task of quantum state generation. After defining the paradigm, we provide two basic lemmas for adiabatic quantum state generation: The Sparse Hamiltonian lemma, which gives a general technique for implementing sparse Hamiltonians efficiently, and, The jagged adiabatic path lemma, which gives conditions for a sequence of Hamiltonians to allow efficient adiabatic state generation. We use our tools to prove that any quantum state which can be gen- erated efficiently in the standard model can also be generated efficiently adiabatically, and vice versa. Finally we show how to apply our techniques to generate superpositions corresponding to limiting distributions of a large class of Markov chains, including the uniform distribution over all perfect * Department of Computer Science and Engineering, Hebrew University, Jerusalem, Israel and mathematical Sci- ences Research Institute, Berkeley, California, e-mail:[email protected] Department of Computer Science, Tel Aviv University, Tel-Aviv, Israel, e-mail:[email protected] 1

Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

Adiabatic Quantum State Generation and Statistical Zero

Knowledge

Dorit Aharonov∗ Amnon Ta-Shma †

Abstract

The design of new quantum algorithms has proven to be an extremelydifficult task. This paper considers a different approach to the problem.We systematically study ’quantum state generation’, namely, which super-positions can be efficiently generated. We first show that all problems inStatistical Zero Knowledge (SZK), a class which contains many languagesthat are natural candidates for BQP, can be reduced to an instance of quan-tum state generation. This was known before for graph isomorphism, butwe give a general recipe for all problems in SZK. We demonstrate the re-duction from the problem to its quantum state generation version for threeexamples: Discrete log, quadratic residuosity and a gap version of closestvector in a lattice.

We then develop tools for quantum state generation. For this task, wedefine the framework of ’adiabatic quantum state generation’ which usesthe language of ground states, spectral gaps and Hamiltonians instead ofthe standard unitary gate language. This language stems from the recentlysuggested adiabatic computation model [20] and seems to be especially tai-lored for the task of quantum state generation. After defining the paradigm,we provide two basic lemmas for adiabatic quantum state generation:

• The Sparse Hamiltonian lemma, which gives a general technique forimplementing sparse Hamiltonians efficiently, and,

• The jagged adiabatic path lemma, which gives conditions for a sequenceof Hamiltonians to allow efficient adiabatic state generation.

We use our tools to prove that any quantum state which can be gen-erated efficiently in the standard model can also be generated efficientlyadiabatically, and vice versa. Finally we show how to apply our techniquesto generate superpositions corresponding to limiting distributions of a largeclass of Markov chains, including the uniform distribution over all perfect

∗Department of Computer Science and Engineering, Hebrew University, Jerusalem, Israel and mathematical Sci-ences Research Institute, Berkeley, California, e-mail:[email protected]

†Department of Computer Science, Tel Aviv University, Tel-Aviv, Israel, e-mail:[email protected]

1

Page 2: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

matchings in a bipartite graph and the set of all grid points inside high di-mensional convex bodies. These final results draw an interesting connectionbetween quantum computation and rapidly mixing Markov chains.

1 Introduction

Quantum computation carries the hope of solving in quantum polynomial time classi-cally intractable tasks. The most notable success so far is Shor’s quantum algorithmfor factoring integers and for finding the discrete log [41]. Following Shor’s algo-rithm, several other algorithms were discovered, such as Hallgren’s algorithm forsolving Pell’s equation [28], Watrous’s algorithms for the group black box model [45],and the Legendre symbol algorithm by Van Dam et al [14]. Except for [14], all ofthese algorithms fall into the framework of the Hidden subgroup problem, and in factuse exactly the same quantum circuitry; The exception, [14], is a different algorithmbut also heavily uses Fourier transforms and exploits the special algebraic structureof the problem. Recently, a beautiful new algorithm by Childs et. al.[10] was found,which gives an exponential speed up over classical algorithms using an entirely dif-ferent approach, namely quantum walks. The algorithm however, works in the blackbox model and solves a fairly contrived problem.

One cannot overstate the importance of developing qualitatively different quan-tum algorithmic techniques and approaches for the development of the field of quan-tum computation. In this paper we attempt to make a step in that direction byapproaching the issue of quantum algorithms from a different point of view.

It has been folklore knowledge for a few years already that the problem of graphisomorphism, which is considered classically hard [33] has an efficient quantum algo-rithm as long as a certain state, namely the superposition of all graphs isomorphicto a given graph,

|αG〉 =∑σ∈Sn

|σ(G)〉 (1)

can be generated efficiently by a quantum Turing machine (for simplicity, we ignorenormalizing constants in the above state and in the rest of the paper). The reasonthat generating |αG〉 suffices is very simple: For two isomorphic graphs, these statesare identical, whereas for two non isomorphic graphs they are orthogonal. A simplecircuit can distinguish between the case of orthogonal states and that of identicalstates, where the main idea is that if the states are orthogonal they will prevent thedifferent states of a qubit attached to them to interfere. One is tempted to assumethat such a state, |αG〉, is easy to construct since the equivalent classical distribution,namely the uniform distribution over all graphs isomorphic to a certain graph, can besampled from efficiently. Indeed, the state |βG〉 =

∑σ∈Sn

|σ〉 ⊗ |σ(G)〉 can be easilygenerated by this argument; However, it is a curious (and disturbing) fact of quantummechanics that though |βG〉 is an easy state to generate, so far no one knows how togenerate |αG〉 efficiently, because we cannot forget the value of |σ〉.

2

Page 3: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

In this paper we systematically study the problem of quantum state generation.We will mostly be interested in a restricted version of state generation, namely gen-erating states corresponding to classical probability distributions, which we looselyrefer to as quantum sampling (or Qsampling) from a distribution. To be more specific,we consider the probability distribution of a circuit, DC , which is the distributionover the outputs of the classical circuit C when its inputs are uniformly distributed.

Denote |C〉 def=

∑z∈0,1m

√DC(z) |z〉. We define the problem of circuit quantum

sampling:

Definition 1. Circuit Quantum Sampling (CQS):

Input: (ε, C) where C is a description of a classical circuit from n to m bits, and0 ≤ ε ≤ 12.

Output: A description of a quantum circuit Q of size poly(|C|) such that |Q(|~0〉)−|C〉 | ≤ ε.

We first show that most of the quantum algorithmic problems considered so farcan be reduced to quantum sampling. Most problems that were considered good can-didates for BQP, such as discrete log (DLOG), quadratic residuosity, approximatingclosest and shortest vectors in a lattice, graph isomorphism and more, belong to thecomplexity class statistical zero knowledge, or SZK (see section 2 for background.)We prove

Theorem 1. Any L ∈ SZK (Statistical Zero Knowledge) can be reduced to a familyof instances of CQS.

The proof relies on a reduction by Sahai and Vadhan [40] from SZK to a completeproblem called statistical difference. Theorem 1 shows that a general solution forquantum sampling would imply SZK ⊆ BQP . We note that there exists an oracleA relative to which SZKA 6⊂ BQPA [1], and so such a proof must be non relativizing.

Theorem 1 translates a zero knowledge proof into an instance of CQS. In general,the reduction can be quite involved, building on the reduction in [40]. Specific exam-ples of special interest turn out to be simpler, e.g., for the case of graph isomorphismdescribed above, the reduction results in a circuit CG that gets as an input a uni-formly random string and outputs a uniformly random graph isomorphic to G. Insection 2 we demonstrate the reduction for three interesting cases: a decision vari-ant of DLOG (based on a zero knowledge proof of Goldreich and Kushilevitz [21]),quadratic residuosity (based on a zero knowledge proof of Goldwasser, Micali andRackoff [24]) and approximating the closest vector problem in lattices (based on azero knowledge proof of Goldreich and Goldwasser [22]). The special cases revealthat although quite often one can look at the zero knowledge proof and directly inferthe required state generation, sometimes it is not obvious such a transition exists atall. Theorem 1, however, tells us such a reduction is always possible.

The problem of what states can be generated efficiently by a quantum computeris thus of critical importance to the understanding of the computational power ofquantum computers. We therefore embark on the task of designing tools for quantum

3

Page 4: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

state generation, and studying which states can be generated efficiently. The recentlysuggested framework of adiabatic quantum computation [20] seems to be tailoredexactly for this purpose, since it is stated in terms of quantum state generation; Letus first explain this framework.

Recall that the time evolution of a quantum system’s state |ψ(t)〉 is described bySchrodinger’s equation:

i~d

dt|ψ(t)〉 = H(t)|ψ(t)〉. (2)

where H(t) is an operator called the Hamiltonian of the system. We will considersystems of n qubits; H is then taken to be local, i.e. a sum of operators, eachoperating on a constant number of qubits. This captures the physical restrictionthat interactions in nature involve only a small number of particles, and means thatthe Hamiltonian H(t) can actually be implemented in the lab. Adiabatic evolutionconcerns the case in which H(t) varies very slowly in time; The qualitative statementof the adiabatic theorem is that if the quantum system is initialized in the groundstate (the eigenstate with lowest eigenvalue) of H(0), and if the modification of Hin time is done slowly enough, namely adiabatically, then the final state will be theground state of the final Hamiltonian H(T ).

Recently, Farhi, Goldstone, Gutmann and Sipser [20] suggested to use adiabaticevolutions to solve NP -hard languages. It was shown in [20, 15] that such adiabaticevolutions can be simulated efficiently on a quantum circuit, and so designing such asuccessful process would imply a quantum efficient algorithm for the problem. Farhiet. al.’s idea was to find the minimum of a given function f as follows: H(0) is chosento be some generic Hamiltonian. H(T ) is chosen to be the problem Hamiltonian,namely a matrix which has the values of f on its diagonal and zero everywhere else.The system is then initialized in the ground state of H(0) and evolves adiabatically onthe convex line H(t) = (1− t

T)H0 + t

THT . By the adiabatic theorem if the evolution

is slow enough, the final state will be the groundstate of H(T ) which is exactly thesought after minimum of f .

The efficiency of these adiabatic algorithms is determined by how slow the adi-abatic evolution needs to be for the adiabatic theorem to hold. It turns out thatthis depends mainly on the spectral gaps of the Hamiltonians H(t). If these spectralgaps are not too small, the modification of the Hamiltonians can be done ’fairly fast’,and the adiabatic algorithm then becomes efficient. The main problem in analyzingthe efficiency of adiabatic algorithms is thus lower bounding the spectral gap; Thisis a very difficult task in general, and hence not much is known analytically aboutadiabatic algorithms. [17, 12, 18] analyze numerically the performance of adiabaticalgorithms on random instances of NP complete problems. It was proven in [15, 39]that Grover’s quadratic speed up [26] can be achieved adiabatically. Lower boundsfor special cases were given in [15]. In [2] it was shown that adiabatic evolutionwith local Hamiltonians is in fact equivalent in computational power to the standardquantum computation model.

In this paper, we propose to use the language of Adiabatic evolutions, Hamilto-

4

Page 5: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

nians, ground states and spectral gaps as a theoretical framework for quantum stategeneration. Our goal is not to replace the quantum circuit model, neither to im-prove on it, but rather to develop a paradigm, or a language, in which quantum stategeneration can be studied conveniently. The advantage in using the Hamiltonianlanguage is that the task of quantum state generation becomes much more natural,since adiabatic evolution is cast in the language of state generation. Furthermore,as we will see, it seems that this language lends itself more easily than the standardcircuit model to developing general tools.

In order to provide a framework for the study of state generation using the adi-abatic language, we define adiabatic quantum state generation as general as we can.Thus, we replace the requirement that the Hamiltonians are on a straight line, withHamiltonians on any general path. Second, we replace the requirement that theHamiltonians are local, with the requirement that they are simulatable, i.e., thatthe unitary matrix e−itH(s) can be approximated by a quantum circuit to within anypolynomial accuracy for any polynomially bounded time t. Thus, we still use thestandard model of quantum circuits in our paradigm. However, our goal is to derivequantum circuits solving the state generation problem, from adiabatic state gener-ation algorithms. Indeed, any adiabatic state generator can be simulated efficientlyby a quantum circuit. We give two proofs of this fact. The first proof follows fromthe adiabatic theorem. The second proof is self contained, and does not requireknowledge of the adiabatic theorem. Instead it uses the simple Zeno effect[38], thusproviding an alternative point of view of adiabatic algorithms using measurements(Such a path was taken also in [11].) This implies that adiabatic state generators canbe used as a framework for designing algorithms for quantum state generation.

We next describe two basic and general tools for designing adiabatic state gener-ators. The first question that one encounters is naturally, what kind of Hamiltonianscan be used. In other words, when is it possible to simulate, or implement, a Hamil-tonian efficiently. To this end we prove the sparse Hamiltonian lemma which givesa very general condition for a Hamiltonian to be simulatable. A Hamiltonian H onn qubits is row-sparse if the number of non-zero entries at each row is polynomiallybounded. H is said to be row-computable if there exists a (quantum or classical) ef-ficient algorithm that given i outputs a list (j,Hi,j) running over all non zero entriesHi,j. As a norm for Hamiltonians we use the spectral norm, i.e. the operator norminduced by the l2 norm on states.

Lemma 1. (The sparse Hamiltonian lemma). IfH is a row-sparse, row-computableHamiltonian on n qubits and ||H|| ≤ poly(n), then H is simulatable.

We note that this general lemma is useful also in two other contexts: first, in thecontext of simulating complicated physical systems on a quantum circuit. Second,for continuous quantum walks [13] which use Hamiltonians. For example, in [10]Hamiltonians are used to derive an exponential quantum speed up using quantumwalks. Our lemma can be used directly to simplify the Hamiltonian implementationused in [10] and to remove the unnecessary constraints (namely coloring of the nodes)which were assumed for the sake of simulating the Hamiltonian.

5

Page 6: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

The next question that one encounters in designing adiabatic quantum state gen-eration algorithms concerns bounding the spectral gap, which as we mentioned beforeis a difficult task. We would like to develop tools to find paths in the Hamiltonianspace such that the spectral gaps are guaranteed to be non negligible, i.e. larger than1/poly(n). Our next lemma provides a way to do this in certain cases. Denote α(H)to be the ground state of H (if unique.)

Lemma 2. (The Jagged Adiabatic Path lemma). Let HjT=poly(n)j=1 be a se-

quence of simulatable Hamiltonians on n qubits, all with polynomially bounded norm,non-negligible spectral gaps and with groundvalues 0, such that the inner product be-tween the unique ground states α(Hj) and α(Hj+1) is non negligible for all j. Thenthere is an efficient quantum algorithm that takes α(H0) to within arbitrarily smalldistance from α(HT ).

To prove this lemma, the naive idea is to use the sequence of Hamiltonians asstepping stones for the adiabatic computation, connecting Hj to Hj+1 by a straightline to create the path H(t). However this way the spectral gaps along the way mightbe small. Instead we use two simple ideas, which we can turn into two more usefultools for manipulating Hamiltonians for adiabatic state generation. The first idea isto replace each Hamiltonian Hj by the Hamiltonian ΠHj

which is the projection onthe subspace orthogonal to the ground states of Hj. We show how to implement theseprojections using Kitaev’s phase estimation algorithm [32]. The second useful idea isto connect by straight lines projections on states with non negligible inner product.We show that the Hamiltonians on such a line are guaranteed to have non negligiblespectral gap. These ideas can be put together to show that the jagged adiabatic pathconnecting the projections ΠHj

is guaranteed to have sufficiently large spectral gap.

We use the above tools to show that

Theorem 2. Any quantum state that can be efficiently generated in the circuit model,can also be efficiently generated by an adiabatic state generation algorithm, and viceversa.

Thus the question of the complexity of quantum state generation is equivalent(up to polynomial terms) in the circuit model and in the adiabatic state generationmodel.

In the final part of the paper we demonstrate how our methods for adiabatic quan-tum state generation work in a particularly interesting domain, namely Qsamplingfrom the limiting distributions of Markov chains. There is an interesting connec-tion between rapidly mixing Markov chains and adiabatic computation. A Markovchain is rapidly mixing if and only if the second eigenvalue gap, namely the differencebetween the largest and second largest eigenvalue of the Markov matrix M , is nonnegligible [4]. This clearly bears resemblance to the adiabatic condition of a nonnegligible spectral gap, and suggests to look at Hamiltonians of the form

HM = I −M. (3)

6

Page 7: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

HM will be a Hamiltonian if M is symmetric; if M is not symmetric but is areversible Markov chain [35] we can still define the Hamiltonian corresponding to it(see section 8.) The sparse Hamiltonian lemma has as an immediate corollary thatfor a special type of Markov chains, which we call strongly samplable, the quantumanalog of the Markov chain can be implemented:

Corollary 1. If M is a strongly samplable Markov chain, then HM is simulatable.

In adiabatic computation one is interested in sequences of Hamiltonians; We thusconsider sequences of strongly samplable Markov chains. There is a particularlyinteresting paradigm in the study of Markov chains where sequences of Markov chainsare repeatedly used: Approximate counting [30]. In approximate counting the ideais to start from a Markov chain on a set that is easy to count, and which is containedin a large set Ω the size of which we want to estimate; One then slowly increasesthe set on which the Markov chain operates so as to finally get to the desired setΩ. This paradigm and modifications of it, in which the Markov chains are modifiedslightly until the desired Markov chain is attained, are a commonly used tool inmany algorithms; A notable example is the recent algorithm for approximating thepermanent [29]. In the last part of the paper we show how to use our techniquesto translate such approximate counting algorithms in order to quantum sample fromthe limiting distributions of the final Markov chain. We show:

Theorem 3. (Loosely:) Let A be an efficient randomized algorithm to approximatelycount a set Ω, possibly with weights; Suppose A uses slowly varying Markov chainsstarting from a simple Markov chain. Then there is an efficient quantum algorithmQ that Qsamples from the final limiting distribution over Ω.

We stress that it is NOT the case that we are interested in a quantum speed upfor sampling from various distributions but rather we are interested in the coherentQsample of the classical distribution.

We exploit this paradigm to Qsample from the set of all perfect matchings ofa bipartite graph, using the recent algorithm by Jerrum, Sinclair and Vigoda [29].Using the same ideas we can also Qsample from all linear extensions of partial or-ders, using Bubley and Dyer algorithm [9], from all lattice points in a convex bodysatisfying certain restrictions using Applegate-Kannan technique [6] and from manymore states. We note that some of these states (perhaps all) can be generated usingstandard techniques which exploit the self reducibility of the problem (see [27]). Westress however that our techniques are qualitatively and significantly different fromprevious techniques for generating quantum states, and in particular do not requireself reducibility. This can be important for extending this approach to other quantumstates.

In this paper we have set the grounds for the general study of the problem ofQsampling and adiabatic quantum state generation, where we have suggested touse the language of Hamiltonians and ground states for quantum state generation.This direction points at very interesting and intriguing connections between quantumcomputation and many different areas: the complexity class SZK and its complete

7

Page 8: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

problem statistical difference [40], the notion of adiabatic evolution [31], the studyof rapidly mixing Markov chains using spectral gaps [35], quantum walks [10], andthe study of ground states and spectral gaps of Hamiltonians in Physics. Hopefully,these connections will point at various techniques from these areas which can be bor-rowed to give more tools for adiabatic quantum state generation; Notably, the studyof spectral gaps of Hamiltonians in physics is a lively area with various recently de-veloped techniques (see [42] and references therein). It seems that a much deeperunderstanding of the adiabatic paradigm is required, in order to solve the most in-teresting open question, namely to design interesting new quantum algorithms. Anopen question which might help in the task is to present known quantum algorithms,eg. Shor’s DLOG algorithm, or the quadratic residuosity algorithm, in the languageof adiabatic computation, in an insightful way.

The rest of the paper is organized as follows. We start with the results related toSZK; We then describe quantum adiabatic computation, define the adiabatic quan-tum state generation framework, and use the adiabatic theorem to prove that anadiabatic state generator implies a state generation algorithm. Next we prove ourtwo main tools: the sparse Hamiltonian lemma, and the jagged adiabatic path lemma.We then use these tools to prove that adiabatic state generation is equivalent to stan-dard quantum state generation. Finally we draw the connection to Markov chainsand demonstrate how to use our techniques to Qsample from approximately count-able sets. In the appendix we give the second proof of transforming adiabatic stategenerators to algorithms using the Zeno effect.

2 Qsampling and SZK

We start with some background about Statistical Zero Knowledge. For an excellentsource on this subject, see Vadhan’s thesis [44] or Sahai and Vadhan [40].

2.1 SZK

A pair Π = (ΠY es,ΠNo) is a promise problem if ΠY es ⊆ 0, 1∗, ΠNo ⊆ 0, 1∗ andΠY es ∩ ΠNo = ∅. We look at ΠY es as the set of all yes instances, ΠNo as the set ofall no instances and we do not care about all other inputs. If every x ∈ 0, 1∗ is inΠY es ∪ ΠNo we call Π a language.

We say a promise problem Π has an interactive proof with soundness error εs andcompleteness error εc if there exists an interactive protocol between a prover P anda verifier V denoted by (P, V ), where V is a probabilistic polynomial time machine,and

• If x ∈ ΠY es V accepts with probability at least 1− εc.

• If x ∈ ΠNo then for every prover P ∗, V accepts with probability at most εs.

When an interactive proof system (Π, V ) for a promise problem Π is run on aninput x, it produces a distribution over ”transcripts” that contain the conversation

8

Page 9: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

between the prover and the verifier. I.e., each possible transcript appears with someprobability (depending on the random coin tosses of the prover and the verifier).

An interactive proof system (Π, V ) for a promise problem Π is said to be ”honestverifier statistical zero knowledge”, and in short HVSZK, if there exists a probabilisticpolynomial time simulator S that for every x ∈ ΠY es produces a distribution ontranscripts that is close (in the `1 distance defined below) to the distribution ontranscripts in the real proof. If the simulated distribution is exactly the correctdistribution, we say the proof system is ”honest verifier perfect zero knowledge, andin short HVPZK.

We stress that the simulator’s output is based on the input alone, and the sim-ulator has no access to the prover. Also, note that we only require the simulatorto produce a good distribution on inputs in ΠY es, and we do not care about otherinputs. This is because for ”No” instances there is no correct proof anyway. We referthe interested reader to Vadhan’s thesis [44] for rigorous definitions and a discussionof their subtleties.

The definition of HVSZK captures exactly the notion of “zero knowledge”; Ifthe honest verifier can simulate the interaction with the prover by himself, in casethe input is in Π, then he does not learn anything from the interaction (exceptfor the knowledge that the input is in Π). We denote by HVSZK the class of allpromise problems that have an interactive proof which satisfies these restrictions.One can wonder whether cheating verifiers can get information from an honest proverby deviating from the protocol. Indeed, in some interactive proofs this happens.However, a general result says that any HVSZK proof can be simulated by one whichdoes not leak much information even with dishonest verifiers [23]. We thus denoteby SZK the class of all promise problems which have interactive proof systems whichare statistically zero knowledge against an honest (or equivalently a general) verifier.

It is known that BPP ⊆ SZK ⊆ AM ∩ coAM and that SZK is closed undercomplement. It follows that SZK does not contain any NP–complete language unlessthe polynomial hierarchy collapses. For this, and other results known about thiselegant class, we refer the reader, again, to Vadhan’s thesis [44].

2.2 The complete problem

Recently, Sahai and Vadhan found a natural complete problem for the class StatisticalZero Knowledge, denoted by SZK. One nice thing about the problem is that it doesnot mention interactive proofs in any explicit or implicit way. We need some factsabout distances between distributions in order to define the problem. For two classicaldistributions p(x), q(x) define their `1 distance and their fidelity (this measure isknown by many other names as well):

|p− q|1 =∑x

|p(x)− q(x)|

F (p, q) =∑x

√p(x)q(x)

9

Page 10: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

We also define the variation distance to be ||p− q|| = 12|p− q|1 so that it is a valuebetween 0 and 1. The following fact is very useful:

Fact 1. (See [37])

1− F (p, q) ≤ ||p− q|| ≤√

1− F (p, q)2

or equivalently

1− ||p− q|| ≤ F (p, q) ≤√

1− ||p− q||2

We can now define the complete problem for SZK:

Definition 2. Statistical Difference (SDα,β)

Input : Two classical circuits C0, C1 with m Boolean outputs.

Promise : ||DC0 −DC1 || ≥ α or ||DC0 −DC1 || ≤ β.

Output : Which of the two possibilities occurs? (yes for the first case and no forthe second)

Sahai and Vadhan [40, 44] show that for any two constants 0 ≤ β < α ≤ 1 suchthat even α2 > β, SDα,β is complete for SZK 1. A well explained exposition can alsobe found in [44].

2.3 Reduction from SZK to Qsampling.

We need a very simple building block.

Claim 1. Let ψ = 1√

2(|0, v〉 + |1, w〉). If we apply a Hadamard gate on the firstqubit and measure it, we get the answer 0 with probability 1 + Real(〈v|w〉)2 and 1with probability 1− Real(〈v|w〉)2.

The proof is a direct calculation. We now proceed to prove Theorem 1.

Proof. Let C0, C1 be an input to SD, C0, C1 are circuits with m outputs. It is enoughto show that SD1/4,3/4 ∈ BQP , given that we can Qsample from the given circuits.Let us first assume that we can Qsample from both circuits with ε = 0 error. Wecan therefore generate the superposition 1

√2(|0〉 |C0〉 + |1〉 |C1〉). We then apply a

Hadamard gate on the first qubit and measure it. We use Claim 1 with v = |C0〉 andw = |C1〉. In our case

〈v|w〉 =∑

z∈0,1m

√DC0(z)DC1(z) = F (DC0 , DC0) (4)

We therefore get 0 with probability 1 + F (DC0 , DC0)2. Thus,1Sahai and Vadhan also show, ([44], Proposition 4.7.1) that any promise problem in HVPZK reduces to SD1/2,0,

where the line above the class denotes complement, i.e., we swap between the yes and no instances.

10

Page 11: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

• If ||DC0 − DC1 || ≥ α, then we measure 0 with probability 1 + F (DC0 , DC0)2 ≤1 +

√1− ||DC0 −DC1 ||22 ≤ 1 +

√1− α22, while,

• If ||DC0 − DC1 || ≤ β, then we measure 0 with probability 1 + F (DC0 , DC0)2 ≥2− ||DC0 −DC1 ||2 ≥ 1− β2.

Setting α = 34 and β = 14 we get that if ||DC0 − DC1 || ≥ α we measure 0 withprobability at most 1 +

√1− α22 ≤ 0.831, while if ||DC0 −DC1 || ≤ β we measure 0

with probability at least 1− β2 ≥ 78 = 0.875. Repeating the experiment O(log(1δ))times, we can decide on the right answer with error probability smaller than δ. If thequantum sampling circuit has a small error (say ε < 1100) then the resulting statesare close to the correct ones and the error introduced can be swallowed by the gap ofthe BQP algorithm.

The above theorem shows that in order to give an efficient quantum algorithmfor any problem in SZK, it is sufficient to find an efficient quantum sampler fromthe corresponding circuits. One can use the theorem to start from a zero knowledgeproof for a certain language, and translate it to a family of circuits which we wouldlike to Qsample from. Sometimes this reduction can be very easy, without the needto go through the complicated reduction of Sahai and Vadhan [40], but in generalwe do not know that the specification of the states is easy to derive. For the sakeof illustration, we give the exact descriptions of the states required to Qsample fromfor three examples, in which the reduction turns out to be much simpler than thegeneral case. These cases are of particular interest for quantum algorithms: discretelog, quadratic residuosity and a gap version of Closest vector in a lattice.

2.4 A promise problem equivalent to Discrete Log

The problem :

Goldreich and Kushilevitz [21] define the promise problem DLPc as:

• Input: A prime p, a generator g of Z∗p and an input y ∈ Z∗

p .

• Promise: The promise is that x = logg(y) is in [1, cp] ∪ [p2 + 1, p2 + cp],

• Output: Whether x ∈ [1, cp] or x ∈ [p2 + 1, p2 + cp]

[21] proves that DLOG is reducible to DLPc for every 0 < c < 1/2. They alsoprove that DLPc has a perfect zero knowledge proof if 0 < c ≤ 1/6. We take c = 1/6and show how to solve DLP1/6 with CQS.

The reduction :

We assume we can solve the construction problem for the circuit Cy,k = Cn,g,y,kthat computes Cy,k(i) = y ·gi( mod p) for i ∈ 0, 1k. The algorithm gets into the

state 1√

2[ |0〉 ∣∣Cgp/2+1,blog(p)c−1

⟩+ |1〉 ∣∣Cy,blog(p)c−3

⟩] and proceeds as in Claim 1.

11

Page 12: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

Correctness :

We have:

∣∣Cgp/2+1,blog(p)c−1

⟩= 1

√2t

2t−1∑i=0

∣∣gp/2+i⟩ (5)

where t is the largest power of 2 smaller than p. Also, as y = gx we have

∣∣Cy,blog(p)c−3

⟩= 1

√2t′

2t′−1∑i=0

∣∣gx+i⟩ (6)

where t′ is the largest power of 2 smaller than p/8. Now, comparing the powersof g in the support of Equations 5 and 6 we see that

• If x ∈ [1, cp] then∣∣Cgp/2+1,blog(p)c−1

⟩and

∣∣Cy,blog(p)c−3

⟩have disjoint supports

and therefore 〈Cy,blog(p)c−3|Cgp/2+1,blog(p)c−1〉| = 0, while,

• If x ∈ [p2+1, p2+cp] then the overlap is large and |〈Cy,blog(p)c−3|Cgp/2+1,blog(p)c−1〉|is a constant.

2.5 Quadratic residuosity

The problem :

we denote xRn if x = y2(modn) for some y, and xNn otherwise. The problemQR is to decide on input x, n whether xRn. An efficient algorithm is known forthe case of n being a prime, and the problem is believed to be hard for n = pqwhere p, q are chosen at random among large primes p and q. A basic fact, thatfollows directly from the Chinese remainder theorem is

Fact 2.

• If the prime factorization of n is n = pe11 pe22 . . . pek

k , then for every x

xRn ⇐⇒ ∀1≤i≤k xRpi

• If the prime factorization of n is n = p1p2 . . . pk then every z ∈ Zn that hasa square root, has the same number of square roots.

We show how to reduce the n = pq case to the CQS (adopting the zero knowledgeproof of [24]).

The reduction : We use the circuit Ca(r) that on input r ∈ Zn outputs z =r2a ( mod n). Suppose we know how to quantum sample Ca for every a. On inputintegers n, x, (n, x) = 1, the algorithm gets into the state 1

√2[|0〉 |C1〉+ |1〉 |Cx〉]

and proceeds as in Claim 1.

12

Page 13: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

Correctness :

We have

|Cx〉 =∑z

√pz |z〉 (7)

where pz = Prr(z = r2x), and

|C1〉 =∑z:zRn

α |z〉 (8)

for some fixed α independent of z.

• If xRn then z = r2x is also a square. Furthermore, as (x, n) = 1 we havepz = Prr(r is a square root of zx) and as every square has the same numberof square roots, we conclude that |Cx〉 = |C1〉 and 〈Cx|C1〉 = 1.

• Suppose xNn. There are only p + q − 1 integers r ∈ Zn that are not co-prime to n. For every r co-prime with n, z = xr2 must be a non-residue(or else xRn as well). We conclude that

∑z:zRn pz ≤ p+ qpq ≈ 0 and so

〈Cx|C1〉 ≈ 0.

We note that for a general n, different elements might have a different number ofsolutions (e.g., try n = 8) and the number of elements not co-prime to n mightbe large, so one has to be more careful.

2.6 Approximating CVP

We describe here the reduction to quantum sampling for a gap problem of CVP(closest vector in a lattice), which builds upon the statistical zero knowledge proofof Goldreich and Goldwasser [22]. A lattice of dimension n is represented by a basis,denoted B, which is an n×n non-singular matrix over R. The lattice L(B) is the setof points L(B) = Bc | c ∈ Z

n, i.e., all integer linear combinations of the columnsof B. The distance d(v1, v2) between two points is the Euclidean distance `2. Thedistance between a point v and a set A is d(v,A) = mina∈A d(v, a). We also denote||S|| the length of the largest vector of the set S. The closest vector problem, CVP,gets as input an n–dimensional lattice B and a target vector v ∈ R

n. The outputshould be the point b ∈ L(B) closest to v. The problem is NP hard. Furthermore, itis NP hard to approximate the distance to the closest vector in the lattice to withinsmall factors, and it is easy to approximate it to within 2εn factor, for every ε > 0.See [22] for a discussion. In [22] an (honest prover) perfect zero knowledge proof forbeing far away from the lattice is given. We now describe the promise problem.

The problem :

• Input: An n–dimensional lattice B, a vector v ∈ Rn and designated distance

d. We set g = g(n) =√nc logn, for some c > 0.

13

Page 14: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

• Promise: Either d(v,L(B)) ≤ d or d(v,L(B) ≥ g · d.• Output: Which possibility happens.

We let Ht denote the sphere of all points in Rn of distance at most t from the

origin.

The reduction : The circuit C0 gets as input a random string, and outputs thevector r + η, where r is a uniformly random point in H2n||B∪v|| ∩ L(B) and ηis a uniformly random point η ∈ Hg2·d. [22] explain how to sample such pointswith almost the right distribution, i.e. they give a description of an efficient suchC0.

We remark that the points cannot be randomly chosen from the real (continu-ous) vector space, due to precision issues, but [22] show that taking a fine enoughdiscrete approximation and a large enough cutoff of the lattice results in an ex-ponentially small error. ¿From now on we work in the continuous world, bearingin mind that in fact everything is implemented in a discrete approximation of it.

Now assume we can quantum sample from the circuit C0. We can then alsoquantum sample from the circuit Cv which we define to be the same circuitexcept that the outputs are shifted by the vector v and become r + η + v. Tosolve the gap problem the algorithm gets into the state 1

√2 [ |0〉 |C0〉+ |1〉 |C1〉 ]

and proceeds as in Claim 1.

Correctness :

If v is far away from the lattice L(B), then the calculation at [22] shows thatthe states |C0〉 and |C1〉 have no overlap and so 〈C0|C1〉 = 0.

On the other hand, suppose v is close to the lattice, d(v,L(B)) ≤ d. Notice thatthe noise η has magnitude about gd, and so the spheres around any lattice pointr and around r+v have a large overlap. Indeed, the argument of [22] shows thatif we express |C0〉 =

∑z pz |z〉 and |C1〉 =

∑z p

′z |z〉 then |p−p′|1 ≤ 1−n−2c. We

see that 〈C0|C1〉 = F (p, p′) ≥ n−2c. Iterating the above poly(n) times we get anRQP algorithm, namely a polynomial quantum algorithm with one sided error.

3 Physics Background

This section gives background required for our definition of adiabatic state genera-tion. We start with some preliminaries regarding the operator norm and the Trotterformula. We then describe the adiabatic theorem, and the model of adiabatic com-putation as defined in [20].

14

Page 15: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

3.1 Spectral Norm

The operator norm of a linear transformation T , induced by the l2 norm is called thespectral norm and is defined by

||T || = maxψ 6=0

|Tψ||ψ|

If T is Hermitian or Unitary (in general, if T is normal, namely commutes with itsadjoint) than ||T || equals the largest absolute value of its eigenvalues. If U is unitary,||U || = 1. Also, ||AB|| ≤ ||A|| · ||B||. Finally, if A = (ai,j) is a D ×D matrix, then||A|| ≤ D2||A||∞ where ||A||∞ = maxi,j |ai,j|.Definition 3. We say a linear transformation T2 α–approximates a linear transfor-mation T1 if ||T1 − T2|| ≤ α, and if this happens we write T2 = T1 + α.

3.2 Trotter Formula

Say H =∑Hm with each Hm Hermitian. Trotter’s formula states that one can

approximate e−itH by slowly interleaving executions of e−tHm . We use the followingvariant of it:

Lemma 3. [37] Let Hi be Hermitian, H =∑M

m=1Hm. Further assume H and Hi

have bounded norm, ||H||, ||Hi|| ≤ Λ. Define

Uδ = [ e−δiH1 · e−δiH2 · . . . · e−δiHM ] · [ e−δiHM · e−δiHM−1 · . . . · e−δiH1 ]

Then ||Uδ − e−2δiH || ≤ O(M · (δΛ)3).

Using the || · || properties stated above we conclude:

Corollary 2. Let Hi be Hermitian, H =∑M

m=1Hm. Assume ||H||, ||Hi|| ≤ Λ. Then,for every t > 0

||U t2δδ − e−itH || ≤ O(t2δ ·M · (δΛ)3) (9)

As ||Uδ − I|| ≤ 2MΛδ we also have ||U bt2δcδ − U t2δ

δ || ≤ 2MΛδ and thus:

Corollary 3. Let Hi be Hermitian, H =∑M

m=1Hm. Assume ||H||, ||Hi|| ≤ Λ. Then,for every t > 0

||U bt2δcδ − e−itH || ≤ O(MΛ · δ +MΛ3t · δ2) (10)

Notice that for every fixed t,M and Λ, the error term goes down to zero with δ.In applications, we will pick δ to be polynomially small, in such a way that the aboveerror term is polynomially small.

15

Page 16: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

3.3 Time Dependent Schrodinger Equation

A quantum state |ψ〉 of a quantum system evolves in time according to Schrodinger’sequation:

i~d

dt|ψ(t)〉 = H(t)|ψ(t)〉 (11)

where H(t) is a Hermitian matrix which is called the Hamiltonian of the physicalsystem. The evolution of the state from time 0 to time T can be described byintegrating Schrodinger’s equation over time. If H is constant and independent oftime, one gets

|ψ(T )〉 = U(T )|ψ(0)〉 = e−iHT |ψ(0)〉 (12)

Since H is Hermitian e−iHT is unitary, and so we get the familiar unitary evolutionfrom quantum circuits. The time evolution is unitary regardless of whether H is timedependent or not.

The groundstate of a HamiltonianH is the eigenstate with the smallest eigenvalue,and we denote it by α(H). The spectral gap of a Hamiltonian H is the differencebetween the smallest and second to smallest eigenvalues, and we denote it by ∆(H).

3.4 The adiabatic Theorem

In the study of adiabatic evolution one is interested in the long time behavior (at largetimes T ) of a quantum system initialized in the ground state of H at time 0 when theHamiltonian of the system, H(t) changes very slowly in time, namely adiabatically.

The qualitative statement of the adiabatic theorem is that if the quantum systemis initialized in the ground state of H(0), the Hamiltonian at time 0, and if themodification of H along the path H(t) in the Hamiltonian space is done infinitelyslowly, then the final state will be the ground state of the final Hamiltonian H(T ).

To make this statement correct, we need to add various conditions and quantifi-cations. Historically, the first and simplest adiabatic theorem was found by Bornand Fock in 1928 [8]. In 1958 Kato [31] improved the statement to essentially thestatement we use in this paper (which we state shortly), and which is usually referredto as the adiabatic theorem. A proof can be found in [36]. For more sophisticatedadiabatic theorems see [7] and references therein.

To state the adiabatic theorem, it is convenient and traditional to work with are-scaled time s = t

Twhere T is the total time. The Schrodinger’s equation restated

in terms of the re-scaled time s then reads

i~d

ds|ψ(s)〉 = T ·H(s)|ψ(s)〉 (13)

where T = dtds

can be referred to as the delay schedule, or the total time.

Theorem 4. (The adiabatic theorem, adapted from [36, 20]). Let H(·) be afunction from [0, 1] to the vector space of Hamiltonians on n qubits. Assume H(·) is

16

Page 17: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

continuous, has a unique ground state, for every s ∈ [0, 1], and is differentiable in allbut possibly finitely many points. Let ε > 0 and assume that the following adiabaticcondition holds for all points s ∈ (0, 1) in which the derivative is defined:

Tε ≥ ‖ ddsH(s)‖

(∆(H(s))2(14)

Then, a quantum system that is initialized at time 0 with the ground state α(H(0))of H(0), and evolves according to the dynamics of the Hamiltonians H(·), ends upat re-scaled time 1 at a state |ψ(1)〉 that is within εc distance from α(H(1)) for someconstant c > 0.

We will refer to equation 14 as the adiabatic condition.

The proof of the adiabatic theorem is rather involved. One way to get intuitionabout it is by observing how the Schrodinger equation behaves when eigenstates areconsidered; If the eigenvalue is λ, the eigenstate evolves by a multiplicative factor eiλt,which rotates in time faster the larger the absolute value of the eigenvalue λ is, andso the ground state rotates the least. The fast rotations are essentially responsible tothe cancellations of the contributions of the vectors with the higher eigenvalues, dueto interference effects.

4 Adiabatic Quantum State Generation

In this section we define our paradigm for quantum state generation, based on theideas of adiabatic quantum computation (and the adiabatic theorem). We would liketo allow as much flexibility as possible in the building blocks. We therefore allowany Hamiltonian which can be implemented efficiently by quantum circuits. We alsoallow using general Hamiltonian paths and not necessarily straight lines. We define:

Definition 4. (Simulatable Hamiltonians). We say a Hamiltonian H on nqubits is simulatable if for every t > 0 and every accuracy 0 < α < 1 the unitarytransformation

U(t) = e−iHt (15)

can be approximated to within α accuracy by a quantum circuit of size poly(n, t, 1/α).

IfH is simulatable, then by definition so is cH for any 0 ≤ c ≤ poly(n). It thereforefollows by Trotter’s equation (3) that any convex combination of two simulatable,bounded norm Hamiltonians is simulatable. Also, If H is simulatable and U is aunitary matrix that can be efficiently applied by a quantum circuit, then UHU † isalso simulatable, because e−itUHU

†= Ue−itHU †.

We note that these rules cannot be applied unboundedly many times in a recursiveway, because the simulation will then blow up. The interested reader is referred to[37, 10] for a more complete set of rules for simulating Hamiltonians.

We now describe an adiabatic path, which is an allowable path in the Hamiltonianspace:

17

Page 18: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

Definition 5. (Adiabatic path). A function H from s ∈ [0, 1] to the vector spaceof Hamiltonians on n qubits, is an adiabatic path if

• H(s) is continuous,

• H(s) is differentiable except for finitely many points,

• ∀s H(s) has a unique groundstate, and

• ∀s, H(s) is simulatable given s.

The adiabatic theorem tells us that the time evolution of a system evolving alongthe adiabatic path will end with the final ground state, if done slowly enough, namelywhen the adiabatic condition holds. Adiabatic quantum state generation is basicallythe process of implementing the Schrodinger’s evolution along an adiabatic path,where we require that the adiabatic condition holds.

Definition 6. (Adiabatic Quantum State Generation). An adiabatic QuantumState Generator (Hx(s), T, ε) has for every x ∈ 0, 1n an adiabatic path Hx(s), suchthat for the given T, ε the adiabatic condition is satisfied for all s ∈ [0, 1] where itis defined. We also require that the generator is explicit, i.e., that there exists apolynomial time quantum machine that

• On input x ∈ 0, 1n outputs α(Hx(0)), the groundstate of Hx(0), and,

• On input x ∈ 0, 1n, s ∈ [0, 1] and δ > 0 outputs a circuit Cx(s) approximatinge−iδHx(s).

We then say the generator adiabatically generates α(Hx(1)).

Remark: We note that in previous papers on adiabatic computation, eg. in [15], adelay schedule τ(s) which is a function of s was used. We chose to work with onesingle delay parameter, T , instead, which might seem restrictive; However, workingwith a single parameter does not restrict the model since more complicated delayschedules can be encoded into the dependence on s.

We will show that every adiabatic quantum state Generator can be efficientlysimulated by a quantum circuit, in Claim 2. We later on prove the other directionof Claim 2, which implies Theorem 2, which shows the equivalence in computationalpower of quantum state generation in the standard and in the adiabatic frameworks.Thus, designing state generation algorithms in the adiabatic paradigm indeed makessense since it can be simulated efficiently on a quantum circuit, and we do not losein computational power by moving to the adiabatic framework and working onlywith ground states. The advantage in working in the adiabatic model is that thelanguage of this paradigm seems more adequate for developing general tools. Afterthe statement and proof of Claim 2, we proceed to prove several such basic tools.Once we develop these tools, we will be able to prove the other direction of theequivalence theorem and apply the tools for generating interesting states.

18

Page 19: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

4.1 Circuit simulation of adiabatic state generation

An adiabatic state generator can be simulated efficiently by a quantum circuit:

Claim 2. (Circuit simulation of adiabatic state generation). Let (Hx(s), T, ε)be an Adiabatic Quantum State Generator. Assume T ≤ poly(n). Then, there existsa quantum circuit that on input x generates the state α(Hx(1)) to within ε accuracy,with size poly(T, 1/ε, n).

Proof. (Based on Adiabatic Theorem) The circuit is built by discretizing timeto sufficiently small intervals of length δ, and then applying the unitary matricese−iH(δj)δ. Intuitively this should work, as the adiabatic theorem tells us that a phys-ical system evolving for time T according to Schrodinger’s equation with the givenadiabatic path will end up in a state close to α(Hx(1)). The formal error analysiscan be done by exactly the same techniques that were used in [15]. We do not givethe details of the proof based on the adiabatic theorem here, since the proof of theadiabatic theorem itself is hard to follow.

We give a second proof of Claim 2. The proof does not require knowledge of theadiabatic theorem. Instead, it relies on the Zeno effect[38], and due to its simplicity,we can give it in full details. We include it in order to give a self contained proof ofClaim 2, and also because we believe it gives a different, illuminating perspective onthe adiabatic evolution from the measurement point of view. We note that anotherapproach toward the connection between adiabatic computation and measurementswas taken in [11]. The full Zeno based proof appears in Appendix A. Here we give asketch.

Proof. (Based on the Zeno effect) As before, we begin at the state α(H(0)), andthe circuit is built by discretizing time to sufficiently small intervals of length δ. Ateach time step j, j = 1, . . . , R, instead of simulating the Hamiltonian as before weapply a measurement determined by H(sj). Specifically, we measure the state in abasis which includes the groundstate α(H(sj)). If R is sufficiently large, the subse-quent Hamiltonians are very close in the spectral norm, and the adiabatic conditionguarantees that their groundstates are very close in the Euclidean norm. Given thatat time step j the state is the groundstate α(H(sj)), the next measurement resultswith very high probability in a projection on the new groundstate α(H(sj+1)). TheZeno effect guarantees that the error probability behaves like 1/R2, i.e. quadraticallyin R (and not linearly), and so the accumulated error after R steps is still small,which implies that the probability that the final state is the groundstate of H(1) isvery high, if R is taken to be large enough.

5 The Sparse Hamiltonian Lemma

Our first concern is which Hamiltonians can be simulated efficiently. We restate thesparse Hamiltonian lemma:

19

Page 20: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

Lemma 1 The sparse Hamiltonian lemma If H is a row-sparse, row-computableHamiltonian on n qubits and ||H|| ≤ poly(n) then H is simulatable.

The main idea of the proof is to explicitly write H as a sum of polynomially manybounded norm Hamiltonians Hm which are all block diagonal (in a combinatorialsense) and such that the size of the blocks in each matrix is at most 2× 2. We thenshow that each Hamiltonian Hm is simulatable and use Trotter’s formula to simulateH .

5.1 The reduction to 2× 2 combinatorially block diagonal matrices.

Let us define:

Definition 7. (Combinatorial block.) Let A be a matrix with rows ROWS(A) andcolumns COLS(A). We say (R,C) ⊆ ROWS(A) × COLS(A) is a combinatorialblock if |R| = |C|, for every c ∈ C, r 6∈ R, A(c, r) = 0, and for every c 6∈ C, r ∈ R,A(c, r) = 0.

A is block diagonal in the combinatorial sense iff there is some renaming of thenodes under which it becomes block diagonal in the usual sense. Equivalently, Ais block diagonal in the combinatorial sense iff there is a decomposition of its rowsinto ROWS(A) =

⋃Bb=1Rb, and of its columns COLS(A) =

⋃Bb=1Cb such that for

every b, (Rb, Cb) is a combinatorial block. We say A is 2 × 2 combinatorially blockdiagonal, if each combinatorial block (Rb, Cb) is at most 2× 2, i.e., for every b either|Rb| = |Cb| = 1 or |Rb| = |Cb| = 2.

Claim 3. (Decomposition lemma). Let H be a row-sparse, row-computable Hamil-tonian over n qubits, with at most D non-zero elements in each row. Then there is a

way to decompose H into H =∑(D+1)2n6

m=1 Hm where:

• Each Hm is a row-sparse, row-computable Hamiltonian over n qubits, and,

• Each Hm is 2× 2 combinatorially block diagonal.

Proof. (Of Claim 3) We color all the entries of H with (D+ 1)2n6 colors. For (i, j) ∈[N ]× [N ] and i < j (i.e., (i, j) is an upper-diagonal entry) we define:

colH(i, j) = (k , i mod k , j mod k , rindexH(i, j) , cindexH(i, j)) (16)

where

• If i = j we set k = 1, otherwise we let k be the first integer in the range [2..n2]such that i 6= j(modk), and we know there must be such a k.

• If Hi,j = 0 we set rindexH(i, j) = 0, otherwise we let rindexH(i, j) be the indexof Hi,j in the list of all non-zero elements in the i’th row of H . cindexH(i, j) issimilar, but with regard to the columns of H .

20

Page 21: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

For lower-diagonal entries, i > j, we define colH(i, j) = colH(j, i). Altogether, we use(n2)3 · (D + 1)2 colors.

For a color m, we define Hm[i, j] = H [i, j] · δcolH(i,j),m, i.e., Hm is H on the entriescolored by m and zero everywhere else. Clearly, H =

∑mHm and each Hm is

Hermitian. Also as H is row-sparse and row-computable, there is a simple poly(n)-time classical algorithm computing the coloring colH(i, j), and so each Hm is alsorow-computable. It is left to show that it is 2× 2 combinatorially block-diagonal.

Indeed, fix a color m. Let us order all the upper-triangular, non-zero elements ofHm in a list NONZEROm = (i, j) | Hm(i, j) 6= 0 and i ≤ j. Say the elements ofNONZEROm are (i1, j1), . . . , (iB, jB). For every element (ib, jb) ∈ NONZEROm

we introduce a block. If ib = jb then we set Rb = Cb = ib while if ib 6= jb then weset Rb = Cb = ib, jb.

Say ib 6= jb (the ib = jb case is similar and simpler). As the color m contains therow-index and column-index of (ib, jb), it must be the case that (ib, jb) is the onlyelement in NONZEROm from that row (or column). Furthermore, as ib mod k 6=jb mod k, and both k, i mod k and j mod k are included in the color m, it must bethe case that there are no elements in NONZEROm that belong to the jb row or ibcolumn (see Figure 1). It follows that (Rb, Cb) is a block. We therefore see that Hm

is 2× 2 combinatorially block-diagonal as desired.

Figure 1: In the upper diagonal side of the matrix Hm: the row and column of (ib, jb) are emptybecause the color m contains the row-index and column index of (i, j), and the jb’th row and ib’thcolumn are empty because m contains k, i mod k, j mod k and i mod k 6= j mod k. The lowerdiagonal side of Hm is just a reflection of the upper diagonal side. It follows that ib, jb is a 2× 2combinatorial block.

Claim 4. For every m, ||Hm|| ≤ ||H||.Proof. Fix an m. Hm is combinatorially block diagonal and therefore its norm ||Hm||is achieved as the norm of one of its blocks. Now, Hm blocks are either

• 1× 1, and then the block is (Hi,i) for some i, and it has norm |Hi,i|, or,

• 2× 2, and then the block is

(0 Ak,`A∗k,` 0

)for some k, `, and has norm |Ak,`|.

It follows that maxm ||Hm|| ≤ maxk,` |Hk,`|. On the other hand ||H|| ≥ maxk,` |Hk,`|.The proof follows.

21

Page 22: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

5.2 2× 2 combinatorially block diagonal matrices are simulatable.

Claim 5. Every 2 × 2 combinatorially block diagonal, row-computable HamiltonianA is simulatable to within arbitrary polynomial approximation.

Proof. Let t > 0 and α > 0 an accuracy parameter.

The circuit :

A is 2 × 2 combinatorially block diagonal. It therefore follows that A’s actionon a given input |k〉 is captured by a 2×2 unitary transformation Uk. Formally,given k, say |k〉 belongs to a 2 × 2 block k, ` in A. We set bk = 2 (for a2 × 2 block) and mink = min(k, `), maxk = max(k, `) (for the subspace towhich k belongs). We then set Ak to be the part of A relevant to this subspace

Ak =

(Amink,mink

Amink,maxk

Amaxk ,minkAmaxk ,maxk

)and Uk = e−itAk . If |k〉 belongs to a 1×1 block

we similarly define bk = 1, mink = maxk = k, Ak = (Ak,k) and Uk = (e−itAk).

Our approximated circuit simulates this behavior. We need two transformations.We define

T1 : |k, 0〉 →∣∣∣bk, mink, maxk, Ak, Uk, k⟩

where Ak is our approximation to the entries of Ak and Uk is our approximation

to e−itAk , and where both matrices are expressed by their four (or one) entries.We use αO(1) accuracy.

Having Uk, mink, maxk, k written down, we can simulate the action of Uk. Wecan therefore have an efficient unitary transformation T2:

T2 :∣∣∣Uk, mink, maxk⟩ |v〉 =

∣∣∣Uk, mink, maxk⟩ ∣∣∣Ukv⟩for |v〉 ∈ Spanmink, maxk.Our algorithm is applying T1 followed by T2 and then T−1

1 for cleanup.

Correctness : Let us denote DIFF = e−itA − T−11 T2T1. Our goal is to show that

||Diff|| ≤ α. We notice that Diff is also 2 × 2 block diagonal, and therefore itsnorm can be achieved by a vector ψ belonging to one of its dimension one or twosubspaces, say to Spanmink, maxk. Let Uk |ψ〉 = α |mink〉 + β |maxk〉 and

Uk |ψ〉 = α |mink〉+ β |maxk〉. We see that:

|ψ, 0〉 T1−→∣∣∣bk, mink, maxk, Ak, Uk, ψ⟩

T2−→∣∣∣bk, mink, maxk, Ak, Uk, Ukψ⟩

= α∣∣∣bk, mink, maxk, Ak, Uk, mink⟩ + β

∣∣∣bk, mink, maxk, Ak, Uk, maxk⟩T−11−→ α |mink, 0〉+ β |maxk, 0〉

22

Page 23: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

where the first equation holds since it holds for |mink〉, |maxk〉 and by linearityit holds for the whole subspace spanned by them. We conclude that |Diff |ψ〉 | =|(Uk− Uk) |ψ〉 | and so ||Diff|| = maxk ||Uk− Uk||. However, by our construction,

||Ak −Ak||∞ ≤ αO(1) and so ||Uk − Uk|| ≤ α as desired.

We proved the claim for matrices with 2 × 2 combinatorial blocks. We remarkthat the same approach works for matrices with m×m combinatorial blocks, as longas m is polynomial in n.

5.3 Proving the sparse Hamiltonian lemma

We now prove the sparse Hamiltonian Lemma:

Proof. (Of Lemma 1.) Let H be row-sparse with D ≤ poly(n) non-zero elements ineach row, and ||H|| = Λ ≤ poly(n). Let t > 0. Our goal is to efficiently simulatee−itH to within α accuracy.

We express H =∑M

m=1Hm as in Claim 3, M ≤ (D+ 1)2n6 ≤ poly(n). We choose∆ such that O(MtΛ3∆2) ≤ α2. Note that 1∆ ≤ poly(t, n) for some large enoughpolynomial. By Claim 5 we can simulate in polynomial time each e−i∆Hm to withinα2Mt/∆ accuracy. We then compute U t2∆

∆ , using our approximations to e−i∆Hm , asin Corollary 3. Corollary 3 assures us that our computation is α close to e−itH , asdesired (suing the fact that for every m, ||Hm|| ≤ ||H|| = Λ ≤ poly(n)). The size ofthe computation is t2∆ · 2M · poly(∆,M, n, α) = poly(n, t, α) as required.

6 The Jagged Adiabatic Path Lemma

Next we consider the question of which paths in the Hamiltonian space guaranteenon negligible spectral gaps. We restate the jagged adiabatic path lemma.

Lemma 2: Let HjT=poly(n)j=1 be a sequence of bounded norm, simulatable Hamiltoni-

ans on n qubits, with non-negligible spectral gaps and with groundvalues 0 such thatthe inner product between the unique ground states α(Hj), α(Hj+1) is non negligiblefor all j. Then there is an efficient quantum algorithm that takes α(H0) to withinarbitrarily small distance from α(HT ).

Proof. (of lemma 2) We replace the sequence Hj with the sequence of HamiltoniansΠHj

where ΠH is a projection on the space orthogonal to the groundstate of Hj,

and we connect two neighboring projections by a line. We prove in claim 6, usingKitaev’s phase estimation algorithm, that the fact that Hj is simulatable impliesthat so is ΠHj

. Also, as projections, ΠHjhave bounded norms, ||ΠHj

|| ≤ 1. It followsthen, by the results mentioned in Section 3, that all the Hamiltonians on the path

23

Page 24: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

H1=(1 00 0)

H2

H3=(1 1

1 -1)

H1H2=(0 00 1)

Figure 2: In the left side of the drawing we see two Hamiltonians H1 and H2 connected by astraight line, and the spectral gaps along that line. In the right side of the drawing we see the sametwo Hamiltonians H1 and H2 connected through a jagged line that goes through a third connectingHamiltonian H3 in the middle, and the spectral gaps along that jagged path. Note that on the leftthe spectral gap becomes zero in the middle, while on the right it is always larger than one.

connecting these projections are simulatable, as convex combinations of simulatableHamiltonians.

We now have to show the Hamiltonians on the path have non negligible spectralgap. By definition ΠHj

has a spectral gap equal to 1. It remains to show, however,that the Hamiltonians on the line connecting ΠHj

and ΠHj+1have large spectral gaps,

which we prove in the simple Claim 7.

We can now apply the adiabatic theorem and get Lemma 2. Indeed, a lineartime parameterization suffices to show that this algorithm satisfies the adiabaticcondition.

We now turn to the proofs of claims 6 and 7.

Claim 6. (Hamiltonian-to-projection lemma). Let H be a Hamiltonian on nqubits such that e−iH can be approximated to within arbitrary polynomial accuracyby a polynomial quantum circuit, and let ‖H‖ ≤ m = poly(n). Let ∆(H) be nonnegligible, and larger than 1/nc, and further assume that the groundvalue of H is 0.Then the projection ΠH , is simulatable.

Proof. We observe that Kitaev’s phase estimation algorithm [32, 37] can be appliedhere to give a good enough approximation of the eigenvalue, and as the spectralgap is non-negligible we can decide with exponentially good confidence whether aneigenstate has the lowest eigenvalue or a larger eigenvalue. We therefore can applythe following algorithm:

• Apply Kitaev’s phase estimation algorithm to write down one bit of informationon an extra qubit: whether an input eigenstate of H is the ground state ororthogonal to it.

24

Page 25: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

• Apply a phase shift of value e−it to this extra qubit, conditioned that it is in thestate |1〉 (if it is |0〉 we do nothing). This conditional phase shift correspondsto applying for time t a Hamiltonian with two eigenspaces, the ground stateand the subspace orthogonal to it, with respective eigenvalues 0 and 1, which isexactly the desired projection.

• Finally, to erase the extra qubit written down, we reverse the first step and un-calculate the information written on that qubit using Kitaev’s phase estimationalgorithm again.

We will also use the following basic but useful claim regarding the convex combi-nation of two projections. For a vector |α〉, the Hamiltonian Hα = I − |α〉〈α| is theprojection onto the subspace orthogonal to α. We prove:

Claim 7. Let |α〉 , |β〉 be two vectors in some subspace, Hα = I − |α〉〈α| and Hβ =I−|β〉〈β|. For any convex combination Hη = (1−η)(I−|α〉〈α|)+η(I−|β〉〈β|, η ∈[0, 1], of the two Hamiltonians Hα, Hβ, ∆(Hη) ≥ |〈α|β〉|.Proof. To prove this, we observe that the problem is two dimensional, write |β〉 =a|α〉 + b|α⊥〉, and write the matrix H in a basis which contains |α〉 and |α⊥〉. Theeigenvalues of this matrix are all 1 except for a two dimensional subspace, where thematrix is exactly (

η|a|2 + (1− η) ηab∗

ηa∗b η|b|2). (17)

Diagonalizing this matrix we find that the spectral gap is exactly√

1− 4(1− η)η|b|2which is minimized for η = 1/2 where it is exactly |a|.

We use the tools we have developed to prove the equivalence of standard andadiabatic state generation complexity, and for generating interesting Markov chainstates. We start with the equivalence result.

7 Equivalence of Standard and Adiabatic State Generation

Theorem 2 asserts that any quantum state that can be efficiently generated in thequantum circuit model, can also be efficiently generated by an adiabatic state gen-eration algorithm, and vice versa. We already saw the direction from adiabatic stategeneration to quantum circuits. To complete the proof of Theorem 2 we now showthe other direction.

Claim 8. let |φ〉 be the final state of a quantum circuit C with M gates, then there is aquantum adiabatic state generator which outputs this state, of complexity poly(n,M).

25

Page 26: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

Proof. W.l.o.g. the circuit starts in the state |0〉. We first modify the circuit so thatthe state does not change too much between subsequent time steps. The reason weneed this will become apparent shortly. To make this modification, let us assumefor concreteness that the quantum circuit C uses only Hadamard gates, Toffoli gatesand Not gates. This set of gates was recently shown to be universal by Shi [43], anda simplified proof can be found in [3] (Our proof works with any universal set withobvious modifications.) We replace each gate g in the circuit by two

√g gates. For√

g we can choose any of the possible square roots arbitrarily, but for concretenesswe notice that Hadamard, Not and Toffoli gates have ±1 eigenvalues, and we choose√

1 = 1 and√−1 = i. We call the modified circuit C ′. Obviously C and C ′ compute

the same function.

The path. We let M ′ be the number of gates in C ′. For integer 0 ≤ j ≤M ′, we set

Hx(jM′) = I − |αx(j)〉〈αx(j)|

where |αx(j)〉 is the state of the system after applying the first j gates of C ′ onthe input x. For s = j+η

M ′ , η ∈ [0, 1), define Hx(s) = (1− η)Hx(j) + ηHx(j + 1).

The spectral gaps are large. Clearly all the Hamiltonians Hx(j) for integer 0 ≤j ≤M ′, have non-negligible spectral gaps, since they are projections. We claimthat for any state β and any gate

√g, |〈β|√g|β〉| ≥ 1√

2. Indeed, represent β as

a1v1 + a2v2 where v1 belongs to the 1-eigenspace of√g and v2 belongs to the

i-eigenspace of√g. We see that |〈β|√g|β〉| = ||a1|2 + i|a2|2|. As |a1|2 + |a2|2 = 1,

a little algebra shows that this quantity is at least 1√2. In particular, setting

β = αx(j) we see that |〈αx(j)|αx(j + 1)〉| ≥ 1√2. It therefore follows by claim 7

that all the Hamiltonians on the line between Hx(j) and Hx(j+1) have spectralgaps larger than 1√

2.

The Hamiltonians are simulatable. Given a state |y〉 we can

• Apply the inverse of the first j gates of C ′,

• If we are in state |x, 0〉 apply a phase shift e−iδ, and

• Apply the first j gates of C ′

which clearly implements e−iδHx(j).

Adiabatic Condition is Satisfied. We have dHds(s0) = limζ→0H(s0 + ζ)−H(s0)ζ.We ignore the finitely many points s = jM ′ where j is an integer in [0,M ′]. Forall other points s, when ζ goes to 0 both H(s0 + ζ) and H(s0) belong to thesame interval. Say they belong to the j’th interval, s0 = j + ηM ′, 0 < η < 1.Then,

H(s0) = (1− η)Hx(j) + ηHx(j + 1)

H(s0 + ζ) = H(j + η +M ′ζM ′) = (1− η −M ′ζ)Hx(j) + (η +M ′ζ)Hx(j + 1)

26

Page 27: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

It follows that H(s0 + ζ)−H(s0) = M ′ζHx(j + 1)−M ′ζHx(j) and dHds(s0) =M ′ · [Hx(j + 1) − Hx(j)]. We conclude that ||dHds|| ≤ 2M ′ and to satisfyEquation (14) we just need to pick T = O(M

′ε

).

8 Quantum State Generation and Markov Chains

Finally, we show how to use our techniques to generate interesting quantum statesrelated to Markov chains.

8.1 Markov chain Background

We will consider Markov chains with states indexed by n bit strings. If M is anergodic (i.e. connected, aperiodic) Markov chain, characterized with the matrix Moperating on probability distributions over the state space, and p is an initial proba-bility distribution, then limt7−→∞ pM t = π where π is called the limiting distributionand is unique and independent of p.

A Markov chain M has eigenvalues between −1 and 1. A Markov chain is saidto be rapidly mixing if starting from any initial distribution, the distribution afterpolynomially many time steps is within ε total variation distance from the limitingdistribution π. [5] shows that a Markov chain is rapidly mixing if and only if itssecond eigenvalue gap is non negligible, namely bounded from below by 1/poly(n).

A Markov chain is reversible if for the limiting distribution π it holds that M [i, j] ·πi = M [j, i]·πj . We note that a symmetric Markov chain M is in particular reversible.Also, for an ergodic, reversible Markov chain M πi > 0 for all i.

In approximate counting algorithms one is interested in sequences of rapidly mix-ing Markov chains, where subsequent Markov chains have quite similar limiting dis-tributions. For more background regarding Markov chains, see [35] and referencestherein; For more background regarding approximate counting algorithms see [30].

8.2 Reversible Markov chains and Hamiltonians

For a reversible M we define

HM = I −Diag(√πi) ·M ·Diag( 1√

πj) (18)

A direct calculation shows that M is reversible iff HM is symmetric. In such a casewe call HM the Hamiltonian corresponding to M . The properties of HM and M arevery much related:

Claim 9. If M is a reversible Markov chain, we have:

• HM is a Hamiltonian with ||HM || ≤ 1.

27

Page 28: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

• The spectral gap of HM equals the second eigenvalue gap of M .

• If π is the limiting distribution of M , then the ground state of HM is α(HM) =

|π〉 def=

∑i

√π(i) |i〉.

Proof. If M is reversible, HM is Hermitian and hence has an eigenvector basis. In

particular I −HM =√

ΠM√

Π−1

and so I −HM and M have the same spectrum. Itfollows that if the eigenvalues of HM are λr then the eigenvalues of M are 1− λr.As a reversible Markov chain, M has norm bounded by 1.

Also, if vr is an eigenvector of HM with eigenvalue λr, then Diag(√π)vr is

the corresponding left eigenvectors of M with eigenvalue 1 − λr. In particular,Diag(

√π)α(HM) = π(M). It therefore follows that α(HM)i =

√πi, or in short

α(HM) = |π〉.This gives a direct connection between Hamiltonians, spectral gaps and ground-

states on one hand, and rapidly mixing reversible Markov chains and limiting distri-bution on the other hand.

8.3 Simulating HM

Not every Hamiltonian corresponding to a reversible Markov chain can be easilysimulated. We will shortly see that the Hamiltonian corresponding to a symmetricMarkov chain is simulatable. For general reversible Markov chains we need somemore restrictions. We define:

Definition 8. A reversible Markov chain is strongly samplable if it is:

• row computable, and,

• Given i, j ∈ Ω, there is an efficient way to approximate πiπj.

Row computability holds in most interesting cases but the second requirement isquite restrictive. Still, we note that it holds in many interesting cases such as allMetropolis algorithms (see [25]). It also trivially holds for symmetric M , where thelimiting distribution is uniform.

As HM [i, j] =√πiπjM [i, j] we see that if M is strongly samplable then HM is

row-computable. As HM has bounded norm, the sparse Hamiltonian lemma implies:

Corollary 1: If a Markov chain M is a strongly samplable then HM is simulatable.

8.4 From Markov chains to Quantum Sampling

We are interested in strongly samplable rapidly mixing Markov chains, so that theHamiltonians are simulatable and have non negligible spectral gaps by claim 9. Toadapt this setting to adiabatic algorithms, and to the setting of the jagged adiabaticpath lemma in particular, we now consider sequences of Markov chains, and define:

28

Page 29: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

Definition 9. (Slowly Varying Markov Chains). Let MtTt=1 be a sequence ofMarkov chains on Ω, |Ω| = N = 2n. Let πt be the limiting distribution of Mt. Wesay the sequence is slowly varying if for all c > 0, for all large enough n, for all1 ≤ t ≤ T |πt − πt+1| ≤ 1− 1/nc.

We prove that we can move from sequences of slowly varying Markov chains toQuantum sampling. We can now state Theorem 3 precisely.

Theorem 3: Let MtTt=1 be a slowly varying sequence of strongly samplable Markovchains which are all rapidly mixing, and let πt be their corresponding limiting dis-tributions. Then if there is an efficient Qsampler for |π0〉, then there is an efficientQsampler for |πT 〉.

Proof. We already saw the Hamiltonians HMt are simulatable and have boundednorm. Also, as the Markov chains in the sequence are rapidly mixing, they have largespectral gaps, and therefore so do the Hamiltonians HMt . To complete the proof weshow that the inner product between the groundstates of subsequent Hamiltonians isnon negligible, and then the theorem follows from the jagged path lemma. Indeed,〈α(HMt)|α(HMt+1)〉 = 〈πt|πt+1〉 =

∑i

√πt(i)πt+1(i) ≥ 1− |πt− πt+1| and therefore is

non-negligible.

Essentially all Markov chains that are used in approximate counting that we areaware of meet the criteria of the theorem. The following is a partial list of stateswe can Qsample from using Theorem 1, where the citations refer to the approximatealgorithms that we use as the basis for the quantum sampling algorithm:

1. Uniform superposition over all perfect matchings of a given bipartite graph [29].

2. All spanning trees of a given graph [9].

3. All lattice points contained in a high dimensional convex body satisfying theconditions of [6].

4. Various Gibbs distribution over rapidly mixing Markov chains using the Metropo-lis filter [35].

5. Log-concave distributions [6].

We note that most if not all of these states can be generated using other simplertechniques. However our techniques do not rely on self reducibility, and are thus qual-itatively different and perhaps extendible in other ways. We illustrate our techniquewith the example of how to Qsample from all perfect matchings in a given bipartitegraph. We also note that if we could relax the second requirement in Definition 8the techniques in this section could have been used to give a quantum algorithm forgraph isomorphism.

29

Page 30: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

8.5 Qsampling from Perfect Matchings

In this subsection we heavily rely on the work of Sinclair, Jerrum and Vigoda [29]who recently showed how to efficiently approximate a permanent of any matrix withnon negative entries, using a sequence of Markov chains on the set of Matchings ofa bipartite graph. The details of this work are far too involved to explain here fully,and we refer the interested reader to [29] for further details.

In a nutshell, the idea in [29] is to apply a Metropolis random walk on the set ofperfect and near perfect matchings (i.e. perfect matchings minus one edge) of thecomplete bipartite graph. Since [29] is interested in a given input bipartite graph,which is a subgraph of the complete graph, they assign weights to the edges suchthat edges that do not participate in the input graph are slowly decreasing until theprobability they appear in the final distribution practically vanishes. The weights ofthe edges are updated using data that is collected from running the Markov chain withthe previous set of weights, in an adaptive way. The final Markov chain with the finalparameters converges to a probability distribution which is essentially concentratedon the perfect and near perfect matchings of the input graph, where the probabilityof the perfect matchings is 1/n times that of the near perfect matching.

It is easy to check that the Markov chains being used in [29] are all stronglysamplable, since they are Metropolis chains. Moreover, the sequence of Markovchains is slowly varying. It remains to see that can quantum sample from the limitingdistribution of the initial chain that is used in [29]. This limiting distribution is adistribution over all perfect and near perfect matchings in the complete bipartitegraph, with each near perfect matching having weight n times that of a perfectmatching, where n is the number of nodes of the given graph. Indeed, to generatethis super-position we do the following:

• We generate∑

π∈Sn|mπ〉, where m in the matching on the bipartite graph in-

duced by π ∈ Sn. We can efficiently generate this state because we can generatea super-position over all permutations in Sn, and there is an easy computationfrom a permutation to a perfect matching in a complete bipartite graph and viceversa.

• We generate the state |0〉 +√n

∑ni=1 |i〉 normalized, on a log(n) dimensional

register. This can be done efficiently because of the low-dimension.

• We apply a transformation that maps |m, i〉 to |0, m〉 when i = 0, and to|0, m− ei〉 for i > 0, where m − ei is the matching m minus the i′th edgein the matching. There is an easy computation from m − ei to m, i and viceversa, and so this transformation can be done efficiently. We are now at thedesired state.

Thus we can apply Theorem 1 to Qsample from the limiting distribution of thefinal Markov chain. We then measure to see whether the matching is perfect or not,and with non negligible probability we project the state onto the uniform distributionover all perfect matchings of the given graph.

30

Page 31: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

9 Acknowledgements

We wish to thanks Umesh Vazirani, Wim van Dam, Zeph Landau, Oded Regev, DaveBacon, Manny Knill, Eddie Farhi, Ashwin Nayak and John Watrous for many inspir-ing discussions. In particular we thank Dave Bacon for an illuminating discussionwhich led to the proof of claim 8.

References

[1] S. Aaronson, Quantum lower bound for the collision problem. STOC 2002, pp.635-642

[2] D. Aharonov, W. van Dam, Z. Landau, S. Lloyd, J. Kempe and O. Regev,Universality of Adiabatic quantum computation, in preparation, 2002

[3] D. Aharonov, A simple proof that Toffoli and Hadamard are quantum universal,in preparation, 2002

[4] N. Alon, Eigenvalues and Expanders. Combinatorica 6(1986), pp. 83-96.

[5] N. Alon and J. Spencer, The Probabilistic Method. John Wiley & Sons, 1991.

[6] D. Applegate and R. Kannan, Sampling and integration of near log-concavefunctions, STOC 1991, pp. 156–163

[7] Y. Avron and A. Elgart, Adiabatic Theorem without a Gap Condition, Commun.Math. Phys. 203, pp. 445-463, 1999

[8] M. Born, V. Fock and B. des Adiabatensatzes. Z. Phys. 51, pp. 165-169, 1928

[9] R. Bubley and M. Dyer, Faster random generation of linear extensions, SODA1998, pp. 350-354.

[10] A. M. Childs, R. Cleve, E. Deotto, E. Farhi, S. Gutmann and D. A. Spielman,Exponential algorithmic speedup by quantum walk, quant-ph/0209131

[11] A. M. Childs, E. Deotto, E. Farhi, J. Goldstone, S. Gutmann, A. J. Landahl,Quantum search by measurement, Phys. Rev. A 66, 032314 (2002)

[12] A. M. Childs, E. Farhi, J. Goldstone and S. Gutmann, Finding cliques by quan-tum adiabatic evolution, quant-ph/0012104.

[13] A. M. Childs, E. Farhi and S. Gutmann, An example of the difference betweenquantum and classical random walks, quant-ph/0103020. Also, E. Farhi and S.Gutmann, Quantum Computation and Decision Trees, quant-ph/9706062

[14] W. van Dam and S. Hallgren, Efficient quantum algorithms for Shifted QuadraticCharacter Problems, quant-ph/0011067

31

Page 32: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

[15] W. van Dam, M. Mosca and U. V. Vazirani, How Powerful is Adiabatic QuantumComputation?. FOCS 2001, pp 279-287

[16] W. van Dam and U. Vazirani, More on the power of adiabatic computation,unpublished, 2001

[17] E. Farhi, J. Goldstone and S. Gutmann, A numerical study of the performanceof a quantum adiabatic evolution algorithm for satisfiability, quant-ph/0007071.

[18] E. Farhi, J. Goldstone, S. Gutmann, J. Lapan, A. Lundgren and D. Preda, Aquantum adiabatic evolution algorithm applied to random instances of an NP-complete problem, Science 292, 472 (2001), quant-ph/0104129.

[19] E. Farhi, J. Goldstone and S. Gutmann, Quantum Adiabatic Evolution Algo-rithms with Different Paths, quant-ph/0208135

[20] E. Farhi, J. Goldstone, S. Gutmann and M. Sipser, Quantum Computation byAdiabatic Evolution, quant-ph/0001106

[21] O. Goldreich and E. Kushilevitz, A Perfect Zero-Knowledge Proof System fora Problem Equivalent to the Discrete Logarithm, Journal of Cryptology, pp.97–116, vol 6 number 2, 1993

[22] O. Goldreich and S. Goldwasser, On the limits of non-approximability of latticeproblems, STOC 1998, pp. 1-9

[23] O. Goldreich, A. Sahai and S. Vadhan, Honest-Verifier Statistical Zero-Knowledge Equals General Statistical Zero-Knowledge, STOC 1998 pp. 399–408,

[24] S. Goldwasser, S. Micali and C. Rackoff, The Knowledge Complexity of Inter-active Proof Systems, SIAM Journal on Computing, SIAM J Comput, Vol 18Number 1, pp. 186–208, 1989

[25] M. Grotschel and L. Lovasz, Combinatorial Optimization: A Survey, Handbookof Combinatorics, North-Holland, 1993

[26] L. Grover, Quantum Mechanics helps in searching for a needle in a haystack,Phys. Rev. Letters, July 14, 1997

[27] L. Grover and T. Rudolph, Creating superpositions that correspond to efficientlyintegrable probability distributions, quant-ph/0208112

[28] S. Hallgren, Polynomial-Time Quantum Algorithms for Pell’s Equation and thePrincipal Ideal Problem, STOC 2002

[29] M. Jerrum and A. Sinclair, E. Vigoda A Polynomial-Time Approximation Algo-rithm for the permanent of a matrix with non-negative entries, STOC 2000

[30] M. Jerrum and A. Sinclair, The Markov chain Monte Carlo method: an approachto approximate counting and integration. in Approximation Algorithms for NP-hard Problems, D.S.Hochbaum ed., PWS Publishing, Boston, 1996.

32

Page 33: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

[31] T. Kato, On the adiabatic theorem of Quantum Mechanics, J. Phys. Soc. Jap.5, pp. 435-439 (1951)

[32] A. Yu. Kitaev, Quantum measurements and the Abelian Stabilizer Problem,quant-ph/9511026

[33] J. Kobler, U. Schoning and J. Turan, The Graph Isomorphism Problem. Birk-jauser, 1993.

[34] Landau and Lifshitz, Quantum Mechanics (Second edition of English Transla-tion), Pergamon press, 1965

[35] L. Lovasz: Random Walks on Graphs: A Survey. Combinatorics, Paul Erdos isEighty, Vol. 2 (ed. D. Miklos, V. T. Sos, T. Szonyi, Janos Bolyai MathematicalSociety, Budapest, 1996, pp. 353–398.

[36] Messiah, Quantum Mechanics, John Willey & Sons (1958)

[37] M. A. Nielsen and I. Chuang, Quantum Computation and Information, Cam-bridge University Press, 2000

[38] See A. Peres, Quantum Theory: Concepts and methods, Kluwer Academic Pub-lishers, 1995

[39] J. Roland and N. Cerf, Quantum Search by Local Adiabatic Evolution Phys.Rev. A 65, 042308 (2002)

[40] A. Sahai and S. P. Vadhan, A Complete Promise Problem for Statistical Zero-Knowledge, FOCS 1997 pp. 448–457

[41] P. W. Shor: Polynomial-Time Algorithms for Prime Factorization and DiscreteLogarithms on a Quantum Computer. SIAM J. Comput. 26(5) 1997, pp. 1484-1509

[42] W. L. Spitzer and S. Starr, Improved bounds on the spectral gap above frustra-tion free ground states of quantum spin chains, math-ph/0212029

[43] Y. Shi, Both Toffoli and Controlled-NOT need little help to do universal quantumcomputation, quant-ph/0205115

[44] S. Vadhan, A Study of Statistical Zero Knowledge Proofs, PhD Thesis, M.I.T.,1999

[45] J. Watrous: Quantum algorithms for solvable groups. STOC 2001, pp. 60-67

A Zeno effect approach to simulating adiabatic Generators

Proof. (Of Claim 2) We concentrate on a time interval [s0, s1], s0 < s1, whereH(·) is continuous on [s0, s1] and differentiable on (s0, s1). We denote ηmax =

33

Page 34: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

maxs∈(s0,s1) ||dHds(s)|| and ∆min = mins∈(s0,s1) ∆(H(s)). We choose R ≥ Θ(η2max∆

2minε).

Notice that R is polynomially related to the schedule time T in the adiabatic condi-tion.

We divide the interval [0, 1] to R time steps. At time step j, j = 1, . . . , R, wemeasure the state with a projective, orthogonal measurement that has the groundstate of H(jR) as one of its answers. We begin at the state α(H(0)).

We need to show our procedure can be implemented efficiently, i.e., that if H issimulatable and has a non negligible spectral gap, then such a measurement can beimplemented efficiently. We also need to show our procedure is accurate, i.e., thatunder the condition of the adiabatic theorem, for the R we have chosen, with veryhigh probability the final state is indeed α(H(1)).

Accuracy :

We first bound the relative change of H(s+ δ) with respect to H(s). For s, s+

δ ∈ [s0, s1], H(s + δ) − H(s) =∫ s+δ

sdHds(s)ds and so ||H(s + δ) − H(s)|| =

|| ∫ s+δs

dHds(s)ds|| ≤ ∫ s+δ

s||dHds(s)||ds ≤ ηmax · δ.

Our next step is to claim that two Hamiltonians that are close to each otherhave close groundstates. This is captured in the following claim, that we provelater.

Claim 10. Let H, J be two Hamiltonians ‖H−J‖ ≤ η. Assume H, J have large

spectral gaps: ∆(H),∆(J) ≥ ∆ Then |〈α(H)|α(J)〉| ≥ 1− 4η2

∆2 .

Having that, we see that since ‖H(j + 1R)−H(jR)‖ ≤ ηmaxR, Claim 10 assertsthat the probability for successful projection at the j′th measurement, i.e. theprobability that the outcome is indeed the groundstate, is 1− O(η2

maxR2∆2

min).The probability we err at any of the R steps is therefore at most O(η2

maxR∆2min)

which is at most ε by our choice of R.

Efficiency :

We use Kitaev’s phase estimation algorithm [32, 37] to give, with polynomiallygood confidence, a polynomially good approximation of the eigenvalue, and wethen measure the eigenvalue. As the spectral gap is non-negligible, this in effectdoes an orthonormal measurement with the eigenstate subspace as one possibleanswer, as desired. The procedure is polynomial because H is simulatable andwe can efficiently approximate e−iHt for every polynomial t.

We finish with the proof of Claim 10.

Proof. (Of Claim 10) W.l.o.g we can assume H and J are positive, otherwise justadd C · I to both matrices, for large enough constant C. This does not effect thespectral norm of the difference, the spectral gaps or the inner product between thegroundstates.

34

Page 35: Adiabatic Quantum State Generation and Statistical Zero Knowledge · 2017-11-22 · Adiabatic Quantum State Generation and Statistical Zero Knowledge Dorit Aharonov Amnon Ta-Shma

Let vi be the H eigenvectors with eigenvalues λ1 < . . . < λN , and ui, µifor J . Again, w.l.o.g, 0 = λ1 ≤ µ1. Notice also that µ1 ≤ λ1 + η, because µ1 =minv:||v||=1 |Jv|, and |Jv1| ≤ |Hv1|+ |(J −H)v1| ≤ λ1 + η.

So, |Jv1| ≤ λ1 + η. On the other hand, express v1 = au1 + bu⊥, with u⊥⊥u1.Then, |Jv1| = |bJu⊥ + aJu1| ≥ |b| · µ2 − |a| · µ1 ≥ |b| · (µ1 + ∆)− |a| · µ1 ≥ |b| ·(λ1 + ∆) − |a| · (λ1 + η). Setting λ1 = 0 we get: η ≥ |b| ·∆ − |a| · η. Let us denotec = ∆η. We see that |a| ≥ c|b| − 1.

We now plug in |a| =√

1− |b|2, and square both sides of the inequality. Weget 1 − |b|2 ≥ 1 − 2c|b| + c2|b|2, i.e., |b| ≤ 2cc2 + 1 ≤ 2cc2 = 2c. Equivalently,

|〈α(H1)|α(H2)〉| = |〈v1|u1〉| = |a| = √1− |b|2 ≥ 1− 4c2 = 1− 4η2∆2 as desired.

35